• Nenhum resultado encontrado

Reply to the comments by Mauz, B. and Antonioli, F. on "Sea Level and Climate Changes during OIS 5e in the Western Mediterranean" Discussion

N/A
N/A
Protected

Academic year: 2023

Share "Reply to the comments by Mauz, B. and Antonioli, F. on "Sea Level and Climate Changes during OIS 5e in the Western Mediterranean" Discussion"

Copied!
9
0
0

Texto

(1)

Discussion

Reply to the comments by Mauz, B. and Antonioli, F. on “Sea Level and Climate Changes during OIS 5e in the Western Mediterranean”

T. Bardají a, J.L. Goy b, C. Zazo c, Cl. Hillaire-Marcel d, C.J. Dabrio e, A. Cabero c, B. Ghaleb d, P.G.

Silva f, J. Lario g

a Departamento de Geología, Edificio Ciencias, Universidad de Alcalá, 28871-Alcalá de Henares, Spain

b Departamento de Geología, Facultad de Ciencias, Universidad de Salamanca, 37008-Salamanca, Spain

c Departamento de Geología, Museo Nacional de Ciencias Naturales, CSIC, José Gutiérrez Abascal, 2, 28006-Madrid, Spain

d Université du Québec à Montréal, GEOTOP-UQAM, Montréal, QC, Canada H3C 3P8

e Departamento de Estratigrafía-UCM and Instituto de Geología Económica-CSIC, Universidad Complutense, 28040-Madrid, Spain

f Departamento de Geología, Universidad de Salamanca, Escuela Politécnica Superior de Avila, 05003-Avila, Spain

g Departamento de Ciencias Analíticas, Facultad de Ciencias, Universidad Nacional de Educación a Distancia, 28040-Madrid, Spain

1. Introduction

We really appreciate the comments by Mauz and Antonioli on our paper (Bardají et al., 2009

geomorphologic maps are often lacking in studies related to the reconstruction of past sea level changes.

In most cases these maps would be the only way of reconstructing past sea level oscillations (see for example,

), but here we respond to the erroneous statements made in their critical reading of our work.

We agree with the authors' historic review of the problems raised when effects of tectonics and sea level oscillations are observed in the record of relative sea-level changes in Western Mediterranean.

We should not forget that the Mediterranean itself constitutes a collision plate boundary, with few stable areas within it. Therefore, the study of the Last Interglacial shorelines cannot be simply analysed according to altitude, as they say. What we do not share with Mauz and Antonioli is that “the only solution to this seemed to be determining the age of a coastal deposit using radiometric, dosimetric and chemical dating techniques”.

It is obvious that we need a chronological approach to be as accurate as possible. First of all, however, an exhaustive field analysis is needed, and this is overlooked by Mauz and Antonioli at least in the Spanish Mediterranean sector. To begin with, field analysis would yield a deeper wider geological knowledge, not only of the precise sections but also on a regional scope. Second, a geomorphologic understanding of the selected sections is necessary.

Geomorphology is the best tool that enables us to understand how many different sea level highstands or intervals we are dealing with in terms of their relation to former and younger sea level indicators (i.e. 3D stratigraphic architecture). Sampling for dating purposes can only be done when the sedimentary sequence is clear, and after detailed geomorphologic mapping. Unfortunately,

Radtke and Schellmann, 2004

2. U-series dating of mollusc shells

), becoming the best way to build robust working hypothesis regarding the age of deposits you are dealing with.

Based on their comments, Mauz and Antonioli pay much more attention to the dating than to field knowledge. Indeed, the major failings found by Mauz and Antonioli in our paper relate first to the U-series dating of mollusc shells. They don't agree with our assumed cementation and climate parameters.

Secondly, our use of the U-series dating technique itself is criticised. Thirdly, the role and appearance of Senegalese fauna in the Mediterranean are questioned. Finally, they also find problems with the oolitic facies and recorded sea level changes.

It is unfortunate that Mauz and Antonioli dedicated so much energy in lengthy comments combining a variety of arguments to explain why they simply do not accept our findings They make use of sentences and phrases extracted “out of context” from published literature (including our own) but curiously miss a key paper.

2.1. U-series data

A few examples with respect to U-series data will illustrate both improper and biased use of extracts.

a. “All but one sample of Zazo et al. (2003), all samples of Hillaire-Marcel et al. (1996) and most samples of Hillaire-Marcel et al. (1986) show initial [234U/238U] activity ratios higher than those expected from a closed system.”

(2)

The 234U/238U ratio does not, per se, indicate the

“closure” of a radioactive system, especially when this system corresponds to the initial incorporation of diagenetic (i.e., post-depositional) uranium in the vadose zone of beach deposits (i.e., with continental U-inputs showing variable 234U-excesses). This is a totally ungrounded argument here. However, we think that the discussion on the correction models for open U-series systems was not the aim of this paper, but was one of the main concerns of a previous paper (Goy et al., 2006

b. Mauz and Antonioli cite

) that should have been consulted by Mauz and Antonioli.

Eggins et al. (2005):

“These results are notable for contradicting established wisdom, which assumes that the most reliable ages can be obtained from the inner layer of mollusc shells due to the arrest of U-uptake following a phase of early U-uptake (e.g. Hillaire-Marcel et al., 1996

It is indeed true but, as explained in many of our papers, any geological setting and diagenetic situation leads to its own specific geochemical characteristics.

Thus, Mauz and Antonioli should also have cited here extracts from other papers of our group illustrating the large array of geochemical behaviour of U-series in beach deposits that we have observed. These range from totally “open system” where late diagenetic, possibly modern U-uptake occurs, to systems rapidly

“closed” following an early diagenetic phase of U- uptake (e.g.,

).”

[Zazo et al., 1999] and et al. (1996) and Labonne and Hillaire-Marcel (2000)

c. “

; two papers where post-depositional U-uptake in mollusc shells are discussed in relation with organic lining decay.

Zazo et al. (2003) ascribe the U-series age of the coral (178 ± 10 ka) to MIS 7.1, but according to the astronomical ages of the marine isotope stages to MIS 6.6. This reveals that (i) the age estimate is inaccurate because the Senegalese fauna was not in the Mediterranean Sea during cool climate periods and (ii) the interpretation of the datum is wrong.”

Well aside from a too hasty linkage of high sea-levels to the SPECMAP timescale (an issue that would require a full paper by itself), and a naive statement

about the improbability of the presence of Senegalese fauna during “cold intervals” in the Mediterranean Sea (which we have never claimed), the authors show some bias in comparing the 178 ± 10 ka datum with the age of a maximum insolation in the Northern Hemisphere. They ignore the fact that all U-series

“measurements” obtained in the cited study were used with a U-flux-correction model to discuss the age of the fossil “coral–uranium” (and not that “of the coral”; see Table 4 and Appendix in Goy et al., 2006

2.2. Climate ).

Aridity in SE Spain is a fact, not an opinion. South- eastern Spain exhibits the most arid climate in Europe, more than any other place in Mediterranean European countries. This is not a freely taken assumption, but it can be referenced in any meteorological or climatic classification of Europe (see for instance Peel et al., 2007

Mauz and Antonioli refer to

). Almeria falls within the Bwh (desert) climate and Murcia in the Bwh (steppic) one, after the Köppen classification, and they experience arid or semiarid climates after the Thornthwaite aridity index.

Tzedakis (2007) to justify their climate assumptions for SE Spain, but although this author makes a wonderful palaeoenvironmental review of the Mediterranean for the last few million years, the references used for SE Spain do not reach the Last Interglacial. Indeed, they all relate to the Holocene–uppermost Pleistocene ([Pons and Reille, 1988] and [PaKaufman et al.

(1996) ntKaufman et al. (1996)aleón-Cano et al., 2003]) or at most to MIS 3 ([Cacho et al., 1999] and [Cacho et al., 2006]). Likewise, the pollen records that Mauz and Antonioli, assume as generalized for the entire Mediterranean during MIS 5e, are from other Mediterranean areas, such as NW Greece, and not from SE Spain. Indeed, Tzedakis (2007) suggests

“specific regional patterns in the character of the interglacial successions, reflecting climate and local intrinsic properties”, specifically talking about the drier locations of southernmost Europe, suggesting that these drier areas have experienced a unique evolutionary history.

(3)

2.3. Cementation

Taking into account these climatic characteristics, which lead to very little leaching of carbonates, and the availability of CaCO3 coming from widespread calcareous substratum, rapid cementation would not be unexpected.

As far as we know, beachrock definitions mention more specifically the morphology and precise location of the deposits rather than their petrography.

They are coastal deposits rapidly cemented by calcium carbonate in the intertidal (foreshore) zone, characterized by the preservation of synsedimentary structures. Cementation of beachrocks can certainly be varied, and is generally characterized by high-Mg micrite or aragonite. However, beachrock cementation is not limited to those types et al. (1996)er, 2007]

3. Time of Senegalese fauna immigration ).

Regarding our MIS 5 deposits, and independently of whether or not they are real beachrocks, the preservation of the prograding bodies within the highly cemented littoral deposits leads us to assume an early stage cementation.

Mauz and Antonioli found great problems assuming that S. bubonius entered into the Mediterranean earlier than MIS 5e. They again go back to the problems found in U-series ages of molluscs, but they overlook the field evidence. In our paper (Bardají et al., 2009) we made an extensive report on the occurrence of S. bubonius in Western Mediterranean MIS 5 levels, and, in our opinion, the exclusive association of S. bubonius to MIS 5e deposits can be ruled out, or at least there should be a reasonable doubt about such an association. This species is more abundant in MIS 5e deposits but these are not the only deposit yielding S. bubonius. On Mallorca Island, for example, S. bubonius appear in a marine terrace deposited during a highstand older (and higher) than MIS 5. Zazo et al. (2003) report a marine unit older than MIS 5 deposits (see Fig. 15 in that paper), located at + 10 m, where Cuerda (1989) and Cuerda and Sacares (1992)

In his turn,

reported the occurrence of Senegalese fauna, including S. bubonius.

Hearty (1987) found Strombus fauna associated with deposits producing aminozone E and mixed E and F–G ratios in Campo de Tiro (Mallorca). These ratios of aminozone F–G are reworked populations in the last Interglacial deposits or indigenous populations in older deposits. He also found Strombus in a unit ascribed to MIS 5.3/5.1 (see Fig. 7 therein). Similarly, Hearty (1986)

Mollusc species can migrate in a larval form, and they develop only if they find suitable temperature, salinity and other ecologic conditions for their survival. In this sense, alkenone unsaturation index from the Alboran Sea (

, in a work described as pioneering by Mauz and Antonioli, stated that “Senegalese (or Strombus) community…is thought to have immigrated into the Mediterranean during the middle and/or late Pleistocene”.

Martrat et al., 2004

4. Sedimentary facies

) show higher than present SST not only during MIS 5 but also during MIS 7.

We are very surprised with the statements that Mauz and Antonioli make about the sedimentary facies of MIS 5 deposits from Mediterranean Spanish coasts.

Sadly, these are incorrect.

We are astonished to read “To our knowledge the last ooid formation at the Iberian Peninsula took place during the last Messinian in the Neogene basins of the Betic mountains (“Terminal Carbonate Complex, Riding et al., 1991

4.1. Wind regime during the Last Interglacial )”, and again when it is said that we find “oolitic sand at the coast adjacent to these Neogene Basins”, suggesting that oolitic facies in MIS 5e deposits from SE Spain are reworked from Neogene deposits. This is simply not true.

We infer a certain wind regime for several reasons, not only from the presence of oolites. First, and the most important point, is that we always find oolitic sediments forming wide dune belts, reaching in some cases heights of more than 20 m (El Altet, Alicante, in Montenat, 1973), where sedimentary structures clearly point to an aeolian origin. The second point is that for oolite formation high energy is needed, and we again cite Wanless and Tedesco (1993) who made

(4)

an extensive report on present day oolite formation in Bahamas. However, Mauz and Antonioli cite the work of Simone (1980)

4.2. Reworking from Neogene oolites

to say that oolites can form both under “calm or agitated water”, but what she actually says is that tangential oolites need, in general, high energy, while radial oolites form preferentially in calm environments.

Assuming as certain the fact that we need some energy for oolite (tangential oolites) formation, the energy requirements can come either from wind- wave agitation or from tides. This is why, together with the extended dune development that does not prevail today, we assume a wind regime different from the present for most of the oolitic settings.

Stronger winds from the east are the only way to explain the precise location of these outcrops, always found in east-facing coastal sectors. Furthermore, the present day wind regime (see Fig. 2B) points to enhanced easterly winds during the warmest season, surprisingly affecting those western Mediterranean sectors where MIS 5e oolites are found.

Here, we emphasize that our palaeoenvironmental interpretations do not come exclusively from petrographic characteristics, but primarily from geomorphologic disposition, sedimentary structures and chronological evolution of sedimentary bodies.

We accept that maybe oolites are not currently forming on the Egypt and Tunisia coasts. We have not studied them and we only make such an affirmation based on literature (Lucas, 1955

References given by Mauz and Antonioli (

).

However, Mauz and Antonioli state that oolites from Tunisia and Alexandria are reworked from MIS 5 and early Holocene deposits, assuming oolite formation during MIS 5 in Tunisia.

[Fabricius et al., 1970], [Strasser et al., 1989] and [El-Sammak and Tucker, 2002]) show the petrographic differences (among others, such as mechanical abrasion, boring by organisms, broken oolite envelopes) between “in situ” and reworked ooids. The percentage of oolite particles within the sediment is also reported as an indicator for “in situ” oolites (Strasser et al., 1989),

who considered 60–80% of oolites a good indicator of “in situ” formation.

Regarding the southeastern Spain oolites, the first regional geological study done in Neogene and Quaternary deposits (Montenat, 1973) already considered the possibility of a reworking origin for the Last Interglacial (then called Tyrrhenian) deposits. Petrographic analyses were conclusive: the Neogene oolites were bigger than Pleistocene ones, with a much thicker cortex, but with only a few preserving their original envelopes; most of them appeared recrystalized and with the nucleus and/or envelopes almost completely dissolved. Pleistocene oolites, on the other hand, preserve their aragonite lamellae and display (under polarized light) the black cross, an anomalous birefringence typical of a tangential arrangement of aragonite crystals. These characteristics therefore eliminate a Neogene origin.

Likewise, petrographic analyses done in samples taken in some of the reported oolitic aeolian dunes always show a percentage of oolites that lies within the limits stated (Strasser et al., 1989

Further, assuming that all MIS 5 oolites are reworked from Neogene deposits implies that all Neogene basins contain oolitic facies of either Terminal Carbonate Complex (Late Messinian) or older units.

This contention is most unfortunate and untrue (see ) as an indicator for “in situ” oolites.

Montenat, 1990

Probably, the most outstanding example of the disconnection between the TCC and MIS 5 is found in the Cope Basin, a tiny basin just north of Cape Cope (Aguilas, Murcia) filled with marine Pliocene to recent sedimentary rocks and having a Betic substratum mainly composed of metamorphic rocks (schists, phyllites, etc). There (see Figs. 4 and 5 in , for a complete review of Neogene basins from SE Spain, or even the available geological maps of Spain, IGME). If we consider some of the most outstanding examples of oolitic dunes from MIS 5, we can see that there is no possibility of such a reworking origin.

Bardají et al., 2009) oolites accumulate in coastal dunes and upper foreshore deposits of MIS 5 age, but no TCC or other oolitic rocks have been found or cited so far. Therefore, no evidence of reworking of

(5)

pre-MIS 5 oolitic deposits can be invoked in this example.

Coated grains and oolites are also found in La Marina (El Pinet, Alicante), here again there are no Neogene oolitic facies directly connected with the coastal sector during MIS 5. The same occurs with the wide dune complexes developed in Callblanque or La Manga del Mar Menor (Murcia), where no oolitic rocks have been found in the Neogene sequences.

In Almeria, aeolian oolitic dunes found just north of Gata Cape (Los Genoveses, San José, La Isleta, etc.) are placed on the large volcanic unit of Gata Cape, where no TCC or other oolitic facies are found. Of course it might be argued that those oolites come from eroded TCC rocks that originally capped the volcanic rocks, such as near Carboneras (Mesa de Roldán), but we must take into consideration that there are no oolitic facies where the Neogene units are still preserved. Even more, the ruggedness of the coastal cliffs makes transport of oolites by longshore currents to small isolated bays extremely improbable.

Finally, reworking from the post-evaporite Messinian oolites outcropping in Sorbas, Almeria–Nijar or Polopos basins also seems physically impossible if we take into consideration several points: the long distance from these basins to these coastal sectors;

lack of geographical connection, now and during last Interglacial times, between Almeria–Nijar basin and eastern Gata Cape littoral, due to the elevations of La Serrata range and the Miocene volcanic outcrops of Gata; absence of oolites anywhere between these basins and the Gata Cape oolite dunes; degree of preservation of oolites in MIS 5 deposits, without any sign of abrasion or boring or broken lamellae; and percentage of oolites.

We sincerely think that prior to making such assertions, Mauz and Antonioli, should have acquired closer and more accurate geological and geomorphological field knowledge of the studied area.

5. Sea level interpretation

Mauz and Antonioli question our data and interpretation of sea level highstands.

In the Campo de Tiro section (in the Balearic Islands) three different littoral deposits from MIS 5.e, defining three different high positions of sea level, have been reported by all the authors that have studied this sequence ([Butzer and Cuerda, 1962], [Hearty, 1987], [Cuerda, 1989], [Goy et al., 1997] and [Zazo et al., 2003]). The ages given by more than 30 samples taken in the same section (Hillaire-Marcel et al., 1996) are in complete agreement not only with other sea level interpretations of the Mallorca littoral (see for instance Hearty, 1987, or Tuccimei et al., 2006) but also for sea level intervals defined for MIS 5.e all over the world (see the review done recently by Hearty et al., 2007). These authors report “significant climatic changes occurred thereafter, midway through and continuing to the end of MIS 5e…”, with a marked oceanographic reorganization and an increase in storminess associated with the end of MIS 5.e. We find the same record of three high sea level indicators for MIS 5e, not only in Mallorca Island but also in other littoral sections from SE Spain (Zazo et al., 2003

Mauz and Antonioli argue that we should have compared our sites with tectonically stable areas in the Mediterranean, providing an example of tectonically stable areas in Sardinia or Sicily. Indeed, Sardinia is reported by

). These indicators record different sea level positions but also they show different sedimentary facies indicating significant climatic changes and an increase in storminess to the end of MIS 5e.

Ferranti et al. (2006) as the reference site for MIS 5.e sea level elevation, arguing that it behaves as a tectonically stable area. We question how an island where MIS 5e markers have been described at elevations ranging from 1.8 to more than 10 m above present sea level can be considered stable? Likewise, several positive sea level oscillations within MIS 5 have been reported in Sardinia (Lecca and Carboni, 2007).

Antonioli et al. (2006) report fossil tidal notches from different sites of the Tyrrhenian coast of Italy at elevations ranging from 3.8 to 10.8 m, stating that

“these elevations are very close to the uncertainty range for the eustatic elevation of MIS 5.e highstand”, and so they conclude that the studied sites are located in almost tectonically stable areas. In

(6)

Sicily (one of the reported sites for tidal notches) there are sea level indicators for MIS 5.e at elevations between 2 and 175 m above mean sea level (Ferranti et al., 2006).

Antonioli et al. (2006) report a steady sea level between 127 and 119 ka, explaining the differences in altitude of tidal notches as due to glacio-hydro- isostatic subsidence, given the fact that the areas reported are “tectonically stable”.

As we see, these authors always refer to MIS 5.e sea level position, but which sea level? The exhaustive review done by Hearty et al. (2007)

There are many references to different highstands, or perhaps it is more accurate to say high sea level oscillations during the last interglacial (MIS 5.e), many of them reported in the previously mentioned work by

report on many places around the world, defining up to six different sea-level intervals for MIS 5, three of them corresponding to high positions of sea level at different heights (+ 2.5 ± 1 m between ca. 132 and 125 ka; above + 3 m between 124 and 121 ka; and between + 6 and + 9 m around 121–119 ka), with at least one “short regression” around 125–124 ka.

Hearty et al. (2007) and many others not reported therein but included in our work (Bardají et al., 2009). Regarding MIS 5.3 or 5.1, the first one is usually reported to have a lower sea level than present, but during MIS 5.1 most authors seem to widely agree about a similar or slightly higher than present sea level ([Hearty et al., 1986], [Hearty and Kaufman, 2000], [Vessica et al., 2000], [Murray- Wallace et al., 2001], [Hearty, 2002], [Muhs et al., 2002], [Potter, 2004], [Wehmiller et al., 2004], [Tuccimei et al., 2006] and [Doar and Kendal, 2008]

Finally, we are well aware of the difficulty of reconstructing a sea level curve in the Mediterranean, and this is why we have not attempted it. We do not infer the global sea level from our sites given in Fig.

7 of our paper (as Mauz and Antonioli suggest). This figure is taken without any modification from the

work by

). So, Mauz and Antonioli should not lightly state that there is no record of sea level above the modern position after 114 ka because there are worldwide examples and references to such a sea level.

Ginés et al. (2001), based on phreatic speleothems in Mallorca Island (Baleares). A more recent and complete sea level curve for Mallorca Island is given by some of the authors in Tuccimei et al. (2006)

Acknowledgements

. However, this curve agrees with our data for the same island and the same sector of the island, both in the number of highstands and in ages. This is why we decided to include it in our work.

Research financed by Projects CGL2008-3998, CGL2008-04000 and GRACCIE-CSD-2007-00067.

CHM acknowledges the financial support from the Science and Engineering Research Council of Canada and the UNESCO Chair for Global Change Study of Université du Québec à Montréal. This paper is a contribution to IGCP495 (Quaternary Land Ocean Interactions: Driving Mechanisms and Coastal Responses), INQUA Commission on Coastal and Marine Processes and CM Work Group Paleoclimatology and Global Change (UCM 910198).

References

Antonioli et al., 2006 F. Antonioli, S. Kershaw and L.

Ferranti, A double MIS 5.5 marine notch?, Quaternary International 145–146 (2006), pp. 19–29.

Bardají et al., 2009 T. Bardají, J.L. Goy, C. Zazo, C.

Hillaire-Marcel, C.J. Dabrio, A. Cabero, B. Ghaleb, P.G. Silva and J. Lario, Sea level and climate changes during OIS 5e in the Western Mediterranean, Geomorphology 104 (2009), pp. 22–37.

Butzer and Cuerda, 1962 K.W. Butzer and J. Cuerda, Coastal stratigraphy of southern Mallorca and its implications for the Pleistocene chronology of the Mediterranean Sea, Journal of Geology 70 (1962), pp. 398–416.

Cacho et al., 1999 I. Cacho, J.O. Grimal, C. Pelejero, M. Canals, F.J. Sierro, J.A. Flore and N. Shackleton, Dansgaard–Oeschger and Heinrich event imprints in Alboran Sea paleotemperatures, Paleoceanography 14 (1999), pp. 698–705

Cacho et al., 2006 I. Cacho, N. Shackleton, H.

Elderfield, F.J. Sierro and J. Grimalt, Glacial rapid

(7)

variability in deep water temperature and δ18O from the western Mediterranean Sea, Quaternary Science Reviews 25 (2006), pp. 3294–3311.

Cuerda, 1989 Cuerda, J., 1989. Los tiempos cuaternarios de las Baleares. Dir. Gral. Cultura.

Conselleria de Cultura, Educació i Esports. Govern Balear, Mallorca, Spain.

Cuerda and Sacares, 1992 Cuerda, J., Sacares, J., 1992. El Quaternari al Migjorn de Mallorca. Dir.

Gral. Cultura, Conselleria de Cultura, Educacio i Esports, Govern Balear. Mallorca, 130 pp.

Doar and Kendal, 2008 W. Doar III and C. Kendal, Late Pleistocene to Holocene coastal marine terraces and sea level curves derived form δ18O proxies: Is the 125 ka high-stand the only higher than present event?, 33rd International Geological Congress, Oslo-2008 (2008) Abstract CD.

Eggins et al., 2005 S.M. Eggins, R. Grün, M.T.

Mcculloch, A.W.G. Pike, J. Chappell, L. Kinsley, G.

Mortimer, M. Shelley, C.V. Murray-Wallace, C.

Spötl and L. Taylor, In situ U-series dating by laser ablation multi-collector ICPMS: new prospects for Quaternary geochronology, Quaternary Science Reviews 24 (2005), pp. 2523–2538.

El-Sammak and Tucker, 2002 A. El-Sammak and M.

Tucker, Ooids from Turkey and Egypt in the Eastern Mediterranean and a love-story of Antony and Cleopatra, Facies 46 (2002), pp. 217–228.

O. Münnich, Early Holocene oöids in modern littoral sands reworked from a coastal terrace, southern Tunisia, Science 169 (1970), pp. 757–760.

Ferranti et al., 2006 L. Ferranti, F. Antonioli, B.

Mauz, A. Amorosi, G. Dai Para, G. Mastronuzzi, C.

Monaco, M. Pappalardo, U. Radtke, P. Renda, P.

Romano, P. Sansó and V. Verrubbi, Markers of the Last Interglacial sea level highstand along the coast of Italy: tectonic implications, Quaternary International 145–146 (2006), pp. 30–54.

Ginés et al., 2001 J. Ginés, J.J. Fornós, A. Ginés, F.

Gracia, C. Delitalia, A. Taddeucci, P. Tuccimei and P.L. Vesica, Els espeleotemes freàtics de les coves litorals de Mallorca: canvis del nivell de la Mediterránia i paleoclima en el Pleistocé superior. In:

G.X. Pons and J.A. Guijarro, Editors, El canvi climàtic: passat, present i futur, Monografias Societat d'Història Natural de les Balears vol. 9 (2001), pp. 33–52.

precipitates. In: D. Nash and S. McLaren, Editors, Geochemical sediments & Landscapes, Blackwell Publishing (2007), pp. 365–390.

Goy et al., 1997 J.L. Goy, C. Zazo and J. Cuerda, Evolución de las áreas margino-litorales de la Costa de Mallorca (I. Baleares) durante el Último y Presente Interglacial Nivel del mar Holoceno y clima, Boletín de Geología y Minería 108 (1997), pp. 127–

135

Goy et al., 2006 J.L. Goy, C. Hillaire-Marcel, C.

Zazo, B. Ghaleb, C.J. Dabrio, A. González, T.

Bardají, J. Civis, M. Preda, A. Yébenes and A.M.

Forte, Further evidence for a relatively high sea level during the penultimate interglacial: open-system U- series ages from La Marina (Alicante, East Spain), Geodinamica Acta 19/6 (2006), pp. 409–426.

Hearty, 1986 P.J. Hearty, An inventory of Last Interglacial (sensu lato) Age Deposits from the Mediterranean Basin: a study of Isoleucine Epimerization and U-series dating, Zeitschrift für Geomorphologie. Supplementband 62 (1986), pp. 51–

69.

Hearty, 1987 P.J. Hearty, New data on the Pleistocene of Mallorca, Quaternary Science Reviews 6 (1987), pp. 245–257.

Hearty, 2002

P.J. Hearty and D.S. Kaufman, Whole-rock aminostratigraphy and Quaternary sea-level history of the Bahamas, Quaternary Research 54 (2000), pp.

163–173.

P.J. Hearty, Revision of the late Pleistocene stratigraphy of Bermuda, Sedimentary Geology 153 (2002), pp. 1–21.

Hearty et al., 1986 P.J. Hearty, G.H. Miller, C.E.

Stearns and B.J. Szabo, Aminostratigraphy of Quaternary shorelines in the Mediterranean basin, Geological Society of America bulletin 97 (1986), pp.

850–858.

Hearty et al., 2007 P.J. Hearty, J.T. Hollin, A.C.

Neumann, M.J. O'Leary and M. McCulloch, Global sea-level fluctuations during the Last Interglaciation

(8)

(MIS 5e), Quaternary Science Reviews 26 (2007), pp.

2090–2112.

Hillaire-Marcel et al., 1986 C. Hillaire-Marcel, O.

Carro, C. Causse, J.L. Goy and C. Zazo, Th/U dating of Stromus bubonius-bearing marine terraces in southeastern Spain, Geology 14 (1986), pp. 613–616.

Hillaire-Marcel et al., 1996 C. Hillaire-Marcel, C.

Garièpy, B. Ghaleb, J.L. Goy, C. Zazo and J. Cuerda, U-series measurements in Tyhrrenian deposits from Mallorca — further evidence for two last-interglacial high sea levels in the Balearic Islands, Quaternary Science Reviews 15 (1996), pp. 53–62. Kaufman et al., 1996 A. Kaufman, B. Ghaleb, J.F. Wehmiller and C. Hillaire-Marcel, Uranium concentration and isotope ratio profiles within Mercenaria shells:

geochronological implications, Geochimica et Cosmochimica Acta 60 (1996), pp. 3735–3746.

Labonne and Hillaire-Marcel, 2000 M. Labonne and C. Hillaire-Marcel, Geochemical gradients within modern and fossil shells of Concholepas concholepas from Northern Chile: an insight into U–Th systematic and diagenetic/authigenic isotope imprints in mollusc shells, Geochimica et Cosmochimica Acta 64 (2000), pp. 1523–1534.

Lecca and Carboni, 2007 L. Lecca and S. Carboni, The Tyrrhenian section of San Giovanni di Sinis (Sardinia): Stratigraphic record o fan irregular single high stand, Rivista Italiana di Paleontología e Stratigrafia 113 (3) (2007), pp. 509–523.

Lucas, 1955 G. Lucas, Oolithes marines actuelles et Calcaires oolithiques récents sur le rivage africain de la Méditerranée Orientale (Égypte et Sud Tunisien), Bulletin Station Océanographique de Salammbô (Tunisie) 52 (1955), pp. 19–38.

Martinson et al., 1987 D.G. Martinson, N.G. Pisias, J.D. Hay, J. Imbrie, T.C. Moore and N.J. Shackleton, Age dating and the original Orbital Theory of the Ice Ages: development of a high-resolution 0 to 300,000 year chronology, Quaternary Research 27 (1987), pp.

1–29

Martrat et al., 2004 B. Martrat, J.O. Grimalt, C.

López-Martínez, I. Cacho, F. Sierro, J.A. Flores, R.

Zahn, M. Canals, J. Curtis and D.A. Hodell, Abrupt temperature change in the Western Mediterranean

over the past 250,000 years, Science 306 (2004), pp.

1762–

Milliman, J.D., Müller, G., Förstner, U. (Eds.), 1974.

Recent Sedimentary Carbonates. Part 1: Marine Carbonates. Ed. Springer, 375 pp.

Montenat, 1973

Montenat, C., (Coord.), 1990. Les Bassins Néogènes du domaine bétique oriental (Espagne) : Tectonique et sédimentation dans un couloir de décrochement.

Doc. et Trav. IGAL (Paris) 12–13, 392 p.

Montenat, C., 1973. Les Formations Néogènes et Quaternaires du levant Espagnol (provinces d'Alicante et de Murcia). Thèse. Paris Orsay. 1170 pp.

Muhs et al., 2002 D.R. Muhs, K.R. Simmons and B.

Steinke, Timing and warmth of the last interglacial period: New U-series evidence from Hawaii and Bermuda and a new fossil compilation for North America, Quaternary Science Reviews 21 (2002), pp.

1355–1383.

Murray-Wallace et al., 2001 C.V. Murray-Wallace, B.P. Brooke, J.H. Cann, A.P. Belpeiro and R.P.

Bourman, Whole-rock aminostratigraphy of the Coorong Coastal Plain, South Australia: towards a one million year record of sea-level highstands, Journal of the Geological Society of London 158 (2001), pp. 111–124.

microbienne dans la cimentation préoce des beachrocks (sédiments intertidaux), Terre and Environment 12 (1998) 131 pp.

Pantaleón-Cano et al., 2003 J. Pantaleón-Cano, E.I.

Yll, R. Pérez Obiol and J.M. Roure, Palynological evidence for vegetational history in semiarid areas of the Western Mediterranean (Amería, Spain), The Holocene 13 (2003), pp. 109–119

Peel et al., 2007 M.C. Peel, B.L. Finlayson and T.A.

Mcmahon, Updated world map of the Köppen-Geiger climate classification, Hydrology and Earth System Science 11 (2007), pp. 1633–1644.

Pons and Reille, 1988 A. Pons and M. Reille, The Holocene and upper Pleistocene pollen record from Padul (Granada, Spain) a new study, Palaeogeography, Palaeoclimatology and Palaeoecology 66 (1988), pp. 243–263

(9)

Potter, 2004 E.K. Potter, MIS 5a and 5c sea-level observations and implications for global ice-melting history, 32 International Geological Congress (Florence) (2004) Abstract Volume 2.

Radtke and Schellmann, 2004 U. Radtke and G.

Schellmann, The Marine Quaternary of Barbados, Geographisches Institut der Universität zu Köln, Germany (2004) 137 pp..

Riding et al., 1991 R. Riding, J.C. Braga and J.M.

Martín, Oolite stromatolites and thrombolites, Miocene, Spain: analogues of recent giant Bahamian examples, Sedimentary Geology 71 (1991), pp. 121–

127

Simone, 1980 L. Simone, Ooids: a review, Earth Science Reviews 16 (1980), pp. 319–355.

Strasser et al., 1989 A. Strasser, E. Davaud and Y.

Jedoui, Carbonate cements in Holocene beachrock:

example from Bahiret el Biban, southeastern Tunisia, Sedimentary Geology 62 (1989), pp. 89–100.

Tuccimei et al., 2006 P. Tuccimei, J. Ginés, M.C.

Delitala, A. Ginés, F. Gràcia, J.J. Fornós and A.

Taddeucci, Last Interglacial sea-level changes in Mallorca Island (Western Mediterranean). High precision U-series data from phreatic overgrowths on speleothems, Zeitschrift fur Geomorphologie 50 (2006), pp. 1–21.

Tzedakis, 2007 P.C. Tzedakis, Seven ambiguities in the Mediterranean palaeoenvironmental narrative, QSR 26 (2007), pp. 2042–2066.

Vessica et al., 2000 P.L. Vessica, P. Tuccimei, B.

Turi, J.J. Fornós, A. Ginés and J. Ginés, Late Pleistocene paleoclimates and sea-level change in the Mediterranean as inferred from stable isotope and U- series studies of overgrowths on speleothems, Mallorca, Spain, Quaternary Science Reviews 19 (2000), pp. 865–879.

Wanless and Tedesco, 1993 H.R. Wanless and L.P.

Tedesco, Comparison of oolitic sand bodies generated by tidal vs. wind-wave agitation. In: B.D.

Keith and C.W. Zuppman, Editors, Mississippian oolites and modern analogs, AAPG Studies in Geology vol. 35, American Association of Petroleum Geologists, Tulsa, OK, USA (1993), pp. 199–225.

Wehmiller et al., 2004 J.F. Wehmiller, K.R.

Simmons, H. Cheng, R.L. Edwards, J. Martin-

McNaughton, L.L. York, D.E. Krantz and C.-C.

Shen, Uranium-series coral age from the US Atlantic Coastal plain. The “80 ka problem” revisited, Quaternary International 120 (2004), pp. 3–14.

Zazo et al., 1999 C. Zazo, P.G. Silva, J.L. Goy, C.

Hillaire-Marcel, B. Ghaleb, J. Lario, T. Bardaji and A. Gonzalez, Coastal uplift in continental collision plate boundaries: data from the Last Interglacial marine terraces of the Gibraltar Strait area (south Spain), Tectonophysics 301 (1999), pp. 95–109.

Zazo et al., 2003 C. Zazo, J.L. Goy, C.J. Dabrio, T. Bardají, C. Hillaire-Marcel, B. Ghaleb, J.-Á. Gonzáles-Delgado and V. Soler, Pleistocene raised marine terraces of the Spanish Mediterranean and Atlantic coasts: records of coastal uplift, sea-level highstands and climate changes, Marine Geology 194 (2003), pp. 103–133.

Referências

Documentos relacionados

The subjects did not have changes in the epithelial IDO expression or in the percentage of IDO positive leukocytes when comparing specimens taken during winter and spring from the same